the authors disclose no potential conflicts of...

38
1 KEAP1 Is a Redox Sensitive Target That Arbitrates the Opposing Radiosensitive Effects of Parthenolide in Normal and Cancer Cells Yong Xu 1 , Fang Fang 2 , Sumitra Miriyala 1 , Peter A. Crooks 3 , Terry D. Oberley 4 , Luksana Chaiswing 4 , Teresa Noel 1 , Aaron K. Holley 1 , Yanming Zhao 1 , Kelley K. Kiningham 5 , Daret K. St. Clair 1† and William H. St. Clair 2† 1 Graduate Center for Toxicology, 2 Department of Radiation Medicine, University of Kentucky, Lexington, KY 40536; 3 Department of Pharmaceutical Sciences, UAMS College of Pharmacy, Little Rock, AR 72205; 4 Department of Pathology, University of Wisconsin, Madison, WI 53705; and 5 Department of Pharmaceutical Sciences, Belmont University School of Pharmacy, Nashville, TN 37212 Running title: Keap1, a central regulator of cellular redox signaling Key words: Keap1, Nrf2, PGAM5, Bcl-xL, parthenolide, radiotherapy, prostate cancer, reactive oxygen species (ROS), redox modification, antioxidant proteins, mitochondrial function Corresponding Authors: William H. St. Clair, M.D., Ph.D., Department of Radiation Medicine, University of Kentucky, College of Medicine, Lexington, KY 40536; Phone: 1-(859) 257-4931; Fax: 1-(859) 323-6486 ; Email: [email protected]. Daret K. St. Clair, Ph.D., Graduate Center for Toxicology, University of Kentucky, College of Medicine, Lexington, KY 40536; Phone: 1- (859) 257-3720; Fax: 1-(859) 323-1059; Email: [email protected]. The authors disclose no potential conflicts of interest on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Upload: others

Post on 30-Oct-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

1

KEAP1 Is a Redox Sensitive Target That Arbitrates the Opposing

Radiosensitive Effects of Parthenolide in Normal and Cancer Cells

Yong Xu1, Fang Fang2, Sumitra Miriyala1, Peter A. Crooks3, Terry D. Oberley4, Luksana

Chaiswing4, Teresa Noel1, Aaron K. Holley1, Yanming Zhao1, Kelley K. Kiningham5, Daret K.

St. Clair1† and William H. St. Clair2†

1Graduate Center for Toxicology, 2Department of Radiation Medicine, University of Kentucky,

Lexington, KY 40536; 3Department of Pharmaceutical Sciences, UAMS College of Pharmacy,

Little Rock, AR 72205; 4Department of Pathology, University of Wisconsin, Madison, WI

53705; and 5Department of Pharmaceutical Sciences, Belmont University School of Pharmacy,

Nashville, TN 37212

Running title: Keap1, a central regulator of cellular redox signaling

Key words: Keap1, Nrf2, PGAM5, Bcl-xL, parthenolide, radiotherapy, prostate cancer, reactive

oxygen species (ROS), redox modification, antioxidant proteins, mitochondrial function

† Corresponding Authors: William H. St. Clair, M.D., Ph.D., Department of Radiation Medicine,

University of Kentucky, College of Medicine, Lexington, KY 40536; Phone: 1-(859) 257-4931;

Fax: 1-(859) 323-6486 ; Email: [email protected]. Daret K. St. Clair, Ph.D., Graduate Center for

Toxicology, University of Kentucky, College of Medicine, Lexington, KY 40536; Phone: 1-

(859) 257-3720; Fax: 1-(859) 323-1059; Email: [email protected].

The authors disclose no potential conflicts of interest

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 2: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

2

Abstract

Elevated oxidative stress is observed more frequently in cancer cells than in normal cells. It

is therefore expected that additional exposure to a low level of reactive oxygen species

(ROS) will push cancer cells toward death, whereas normal cells might maintain redox

homeostasis through adaptive antioxidant responses. We previously demonstrated that

parthenolide enhances ROS production in prostate cancer cells through activation of

NADPH oxidase. The present study identifies KEAP1 as the downstream redox target that

contributes to parthenolide's radiosensitization effect in prostate cancer cells. In vivo,

parthenolide increases radiosensitivity of mouse xenograft tumors but protects normal

prostate and bladder tissues against radiation-induced injury. Mechanistically,

parthenolide increases the level of cellular ROS and causes oxidation of thioredoxin (TrX)

in prostate cancer cells, leading to a TrX-dependent increase in a reduced state of KEAP1,

which in turn leads to KEAP1-mediated PGAM5 and Bcl-xL (BCL2L1) degradation. In

contrast, parthenolide increases oxidation of KEAP1 in normal prostate epithelial cells,

leading to increased Nrf2 (NFE2L2) levels and subsequent Nrf2-dependent expression of

antioxidant enzymes. These results reveal a novel redox-mediated modification of KEAP1

in controlling the differential effect of parthenolide on tumor and normal cell

radiosensitivity. Further, they show it is possible to develop a tumor-specific

radiosensitizing agent with radioprotective properties in normal cells.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 3: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

3

Introduction

It is well-documented that cancer cells are usually under more oxidative stress than normal cells

are, in part due to a hyperactive metabolism that fuels their rapid growth (1, 2). Thus, a therapy

designed to increase reactive oxygen species (ROS) to a level above the threshold for cancer cell

death, but to an adaptable level for normal cells, would be an attractive strategy to selectively kill

cancer cells (3, 4). Redox homeostasis is thought to regulate many cellular processes that are

essential for maintenance of normal physiological conditions but is aberrantly modulated in

cancers (5, 6). The functions of ROS are both beneficial and deleterious due to their dual role in

the prosurvival and antisurvival pathways. As a secondary messenger in cell signaling, ROS are

required for normal development and can initiate adaptive responses in cellular defense (7, 8).

On the other hand, ROS cause structural damage and functional decline in DNA, proteins and

lipids, and consequently act as an anti-tumorigenic factor by inducing cell senescence and

apoptosis (9, 10). Indeed, ROS-mediated cell death is an important basis for radiotherapy and

many chemotherapeutic treatments (11, 12). Currently, these therapeutic strategies are being

used to kill cancer cells without benefit of a rational design that exploits the intrinsic differences

in the cellular redox status of normal cells and cancer cells.

Antioxidant defense systems are essential for the regulation of ROS levels, which have an

important function in the maintenance of cellular redox hemostasis. Mounting evidence

demonstrates that a decline in antioxidant function may be involved in tumorigenesis due to

prooxidant conditions that result from ROS accumulation. For example, manganese superoxide

dismutase (MnSOD) is down-regulated in many types of cancer, and overexpression of MnSOD

results in suppression of tumorigenesis (13-15). High levels of antioxidants caused by therapy-

mediated activation of pro-survival pathways, such as NF-κB and Nrf2, are thought to protect

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 4: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

4

cancer cells against treatment (16-18). Thus, inhibition of prosurvival pathways has been a

traditional strategy to enhance therapeutic efficacy.

Increasing evidence demonstrates that certain mild prooxidant compounds derived from

natural herbal medicines might enhance some anticancer treatments by modulating the redox

state of cancer cells to high prooxidant levels (19, 20). Parthenolide, an active ingredient derived

from the traditional anti-inflammatory medical plant feverfew (Tanacetum parthenium), belongs

to the family of sesquiterpene lactones containing an α-methylene-γ-lactone moiety and an

epoxide group, which is able to conjugate thiols of proteins through a Michael addition reaction

(21). In addition to its anti-inflammatory effect, parthenolide appears to be toxic to a variety of

cancer cells (22-25). Importantly, parthenolide has no cytotoxic effect on normal cells (25).

Mechanistically, parthenolide has been shown to increase apoptosis in cancer cells through

inhibition of multiple prosurvival pathways, such as NF-κB and PI3K-AKT (19, 26). However,

these findings do not explain why parthenolide is not toxic to normal cells.

Post-translational modification is a key mechanism by which proteins dramatically

increase their functional diversity. Reversible redox modification of protein cysteine residues

plays an important role in vital cell signaling pathways related to many physiological and

pathogenic processes (27, 28). The Keap1-Nrf2 pathway is one of the main cellular defense

mechanisms against oxidative stress (29, 30). The present study examines the role of cysteine

modifications in modulating radiation responses in prostate cancer cells versus normal prostate

epithelial cells. It elucidates the functional link between redox modulation and cell signaling

transduction pathways, and it provides evidence for the differential effect of parthenolide on

cellular redox status in normal and cancer cells. The results reveal that parthenolide-mediated

redox modification of Keap1 serves as a central regulator of differential responses to radiation

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 5: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

5

therapy in normal and tumor cells. The present study provides a proof-of-concept for utilizing an

intrinsic difference in cellular redox conditions to kill tumor cells while protecting normal cells

from the unwanted side effects of radiation.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 6: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

6

Materials and Methods Cell culture, cell transfection, treatment, and cell survival analysis. Human prostate

carcinoma/adenocarcinoma LNCaP, PC3 and DU145 cell lines as well as human prostate

epithelial viral transformed PZ-HPV-7 (PZ) and RWPE-1 cell lines were obtained from

American Type Culture Collection (ATCC). Normal prostate epithelial PrEC cells were

purchased from Lonza. All cell lines were cultured and maintained in the media recommended in

the manufacturer’s protocols. Plasmid cloned Keap1 cDNA and Bcl-xL cDNA (OriGene) and

siRNAs for knocking down Keap1, Nrf2, thioredoxin (TrX), PGAM5, and Bcl-xL (Dharmacon)

were transfected into cultured cells prior to treatment. Parthenolide and its water soluble prodrug,

dimethylamino-parthenolide (DMAPT), were synthesized as previously described (31). The cells

were treated with 0-5 µM parthenolide followed by irradiation by a 250 kV X-ray machine

(Faxitron X-ray Corp.) with peak energy of 130 kV, 0.05 mm Al filter, at a dose of 0 to 6 Gy.

Cell survival rates were quantified by colony survival fraction, Trypan blue exclusion assay, and

MTT assay, as previously described (25, 32).

Animals. Four- to five-week-old male NCRNU (nu/nu athymic nude) mice were obtained from

Taconic (Hudson). For formation of xenograft tumors, 1.8 x 106 cells mixed in Matrigel (BD

Biosciences) were subcutaneously injected into the right flank of the mice. Tumor volumes were

routinely measured and their sizes calculated based on a protocol described elsewhere (33).

Animals with an average tumor size of 500 mm3 were randomized into several groups for

DMAPT and radiation treatments. The tumors were treated five times with 10 mg/kg DMAPT

and 3 Gy IR. The tumor tissues were collected, and 100 µg of tissue were lysed to quantify

amounts of oxidized or reduced Keap1 and levels of downstream proteins. To determine the

protective effect of parthenolide against radiation damage, mice were pretreated with DMAPT

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 7: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

7

followed by radiation treatment (5 x 3Gy). Prostate and bladder tissues were fixed, embedded

and processed for routine Electron microscopy (EM). The embedded blocks were sectioned and

transferred to copper grids and counterstained with uranyl acetate, followed by lead citrate. Grids

were observed in an electron microscope (Hitachi H-600) operated at 75 kV. Mitochondrial

damage was quantified by a pathologist (TDO) using identified ultrastructural changes including

swelling, vacuolization, myelination, disorganization and loss of cristae, lysosomal degradation,

and membrane disruption. Animal experimental procedures were approved by the Institutional

Animal Care and Use Committee of the University of Kentucky, Approval Protocol No.

01077M2006.

Quantification of intracellular superoxide and prooxidant. Dihydroethidium (DHE,

Invitrogen), which exhibits blue fluorescence in the cytosol until oxidized, was used to estimate

the levels of superoxide after parthenolide treatment. To confirm the level of superoxide induced

by parthenolide, the cells were pretreated with 50 μg PEG-SOD (Sigma) for 1 h followed by

parthenolide treatment. Antimycin (Sigma) was used as a positive control because it has been

shown to increase superoxide in all tested cell lines. A dichlorofluorescein (DCF) assay was used

to quantify the levels of intracellular ROS after parthenolide treatment. The cells were labeled by

both carboxy-H2DCFDA (sensitive to oxidation, Invitrogen) and carboxy-DCFDA (insensitive to

oxidation, Invitrogen). The H2DCFDA:DCFDA ratio was used to optimize the controls of cell

number, dry uptake, and ester cleavage. The procedures for both DHE and DCF were performed

by the University of Kentucky Flow Cytometer Facility using a FACScan protocol provided by

Dr. Douglas R. Spitz (34).

Measurement of oxygen consumption rates (OCR). To determine how parthenolide changes

mitochondrial function in cancer and normal cells, a Seahorse Bioscience XF24 Extracellular

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 8: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

8

Flux Analyzer was used to measure OCR after parthenolide treatment. The XF24 creates a

transient, 7-μl chamber in specialized microplates, which allows determination of oxygen and

proton concentrations in real time. To allow comparison between experiments, data are

expressed as the OCR in pmol/min or the rate of extracellular acidification in mpH/min. Reserve

capacity, an important index that indicates capacity of mitochondrial respiration, is calculated by

subtracting baseline OCR from maximal OCR.

Detection of oxidized and reduced forms of Keap1 protein. 3-N-maleimido-propionyl

biocytin was used to selectively label sulfhydryl (SH) and then was detected by biotin-

streptavidin integration on the blots, as previously described (35). To quantify disulfide (S-S)

bonds, the SH form was stabilized by treating with N-ethylmaleimide and then the S-S bonds

were reduced by treating with 2-mercaptoethanol. To identify SH and S-S moieties of Keap1

protein, the labeled proteins were immunoprecipitated by Keap1 antibody (abcam) and subjected

to SDS-PAGE, followed by detection with horseradish peroxidase-conjugated streptavidin

(Sigma).

Immunoblots and immunoprecipitation. Homogenized cells and tumor tissues were

electrophoresed on an 8% (w/v) SDS-PAGE gel, transferred onto a nitrocellulose membrane, and

subsequently incubated with primary antibodies against Keap1 (abcam), Nrf2 (abcam), MnSOD

(Upstate Biotech), CuZnSOD (eBiosci), Gpx (abcam), catalase (Millipore), TrX (BD Sciences),

PGAM5 (Santa Cruz Biotech.), Bcl-xL (Santa Cruz Biotech.), LC3B (cell signaling), β-actin

(sigma), and pCNA (Santa Cruz Biotech.). All secondary antibodies were obtained from Santa

Cruz Biotech. Immunoblots were visualized using an enhanced chemiluminescence detection

system (Amersham Pharmacia Biotech.). For immunoprecipitation, cell extracts were incubated

overnight with one primary antibody at 4ºC and integrated with a protein A/G agarose (Santa

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 9: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

9

Cruz Biotech). Immunocomplexes were precipitated and fractionated on an SDS-PAGE gel.

Interacting proteins were detected by immunoblotting with their primary antibodies.

SOD enzymatic assay. MnSOD activities were measured by the nitroblue tetrazolium (NBT)-

bathocuproin sulfonate (BCS) reduction inhibition method. Sodium cyanide (2 mM) was used to

inhibit CuZnSOD activity (36).

Statistical data analyses. Multiple independent experiments were performed for each set of data

presented. Images in Immunoblots were quantified using Carestream Molecular Imaging

software (Carestream Health Inc.). Statistical significance was analyzed using one-way ANOVA

and Tukey's Multiple Comparison Test, followed by data analysis with Graphpad Prism.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 10: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

10

Results

Parthenolide enhances radiosensitivity of prostate tumors but protects normal tissues from

radiation. Our previous studies demonstrated that parthenolide is able to selectively enhance the

radiosensitivity of prostate cancer cells without injury to normal prostate epithelial cells (24, 25).

To determine whether parthenolide enhances radiotherapy in vivo, human prostate cancer PC3

cells were subcutaneously implanted into the right flank of nude male mice. When tumor size

reached a volume of 500 mm3, animals were randomized into groups according to treatment

consisting of saline or 10 mg/kg DMAPT. One hour after saline or DMAPT was administered,

the tumors were treated with fractionated radiation of 1 Gy or 2 Gy per day for 5 days, followed

by routine measurement of tumor volume. The tumor growth curves are shown in Fig. 1A. Mice

were humanely killed when a tumor reached the maximum size of 2000 mm3. Tumor growth was

clearly delayed in the treatment groups, particularly when the drug and radiation were combined,

compared to growth in the untreated group. The tumor growth rates after treatment were

compared according to the days needed for tumor volume to reach 2000 mm3. DMAPT

significantly enhanced radiotherapeutic efficiency compared to the effects of radiation treatment

alone (Fig. 1B). A separate group of nude male mice that had no cancer cell implantation was

treated with DMAPT and radiation to determine the toxicity of DMAPT to organs that can be

affected by radiation therapy of prostate cancer. Prostates and bladders of the animals were

examined by light and EM. At 60 days after irradiation, no gross pathology was observed (data

not shown). However, ultrastructural damage was clearly observed by EM (Fig. 1C).

Mitochondrial damage was most pronounced and this was morphometrically analyzed. The

number of damaged mitochondria in prostate and bladder was proportional to radiation exposure.

Pretreatment with DMAPT significantly reduced the number of damaged mitochondria in both

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 11: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

11

organs compared to the group without DMAPT treatment (Fig. 1C and D). These results indicate

that DMAPT, the water soluble prodrug of parthenolide, is a promising agent for selectively

enhancing the sensitivity of prostate cancer cells to radiation while protecting normal tissues

from damage caused by radiation.

Parthenolide differentially modulates cellular ROS levels in cancer and normal cells. To

determine the effect of parthenolide on redox homeostasis in cancer and normal cells, the levels

of superoxide and total ROS after parthenolide treatment were measured using flow cytometry.

The mean fluorescence intensity of DHE and the H2DCFDA:DCFDA ratio were higher in

prostate cancer PC3 cells than in normal prostate PZ and PrEC cells, indicating higher basal

levels of superoxide and total ROS, respectively (Fig. 2A-B). Following parthenolide treatment,

DHE and DCF levels increased further in PC3 cells but declined slightly in both types of non-

cancerous cells. Combining PEG-SOD with parthenolide treatment restored the basal level of

DHE fluorescence. As positive controls, addition of equivalent concentrations of the ROS

stimulating agents antimycin (Fig. 2A) and PMA (Fig. 2B) caused high levels of ROS in all the

tested cell lines.

The levels of antioxidant proteins were also quantified (Fig. 2C). Parthenolide altered the

protein level of the antioxidant enzymes, in particular mitochondria-localized antioxidant

enzymes. MnSOD and glutathione peroxidase (GpX) were significantly reduced in all three

parthenolide-treated prostate cancer cell lines. Intriguingly, parthenolide had the opposite effect

in the three normal prostate cell lines. However, neither cancer nor normal cells showed any

obvious changes after parthenolide treatment for the major cytosolic superoxide removal protein,

copper and zinc-containing SOD (CuZnSOD). This observation was confirmed by quantification

of the corresponding enzyme activity (Fig. 2D). These results suggest that parthenolide-mediated

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 12: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

12

alteration of cellular redox status is mediated, at least in part, by changing the activities of

antioxidant enzymes in mitochondria.

To probe whether altering cellular redox status is associated with a change in

mitochondrial respiration, the OCR in the parthenolide-treated cells was measured using a

Seahorse Bioscience FX OxygenFlux Analyzer. The basal and maximal OCR in normal cells

was higher than in cancer cells (Fig. 2E). Importantly, parthenolide was able to increase the OCR

and reserve capacity in PZ cells, whereas parthenolide had no effect on PC3 OCR. Finally, the

cytotoxicity of parthenolide was tested in all the cell lines using an MTT assay, which requires

active mitochondria. As shown in Fig. 2F, parthenolide was toxic to all the cancer cells but not to

the normal cell lines. Taken together, these results suggest that changes in cellular redox status

and mitochondrial function may be a cause for the differential biological effects of parthenolide

on cancer and normal cells.

Keap1 is susceptible to parthenolide-mediated redox modification. Keap1, a redox sensitive

protein, has been reported to play an important role in cell survival under oxidative stress (29).

To investigate whether parthenolide modifies Keap1 function, a Keap1 antibody linked to biotin

was used to immunoprecipitate redox-modified Keap1 protein, and the presence of oxidized (-S-

S-) and reduced (-SH) cysteine residues was detected using a secondary antibody linked to

streptavidin. In the three normal cell lines, parthenolide increased the oxidized form of Keap1

but decreased the reduced form of Keap1 (Fig. 3A). Interestingly, the results from the three

cancer cell lines appeared to be completely opposite to results observed in normal cells treated

with parthenolide: the level of the oxidized form was decreased, but the level of the reduced

form was increased (Fig. 3B). To verify that the observed increase in reduced Keap1 also

occurred in vivo, mouse xenograft tumor tissues with and without DMAPT treatment were also

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 13: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

13

used for determination of Keap1 redox status. Consistent with data obtained from cultured tumor

cells treated with parthenolide, the in vivo results show that parthenolide decreased the oxidized

form of Keap1 but increased the reduced form of Keap1 in the tumors (Fig. 3C). Changes in

antioxidant proteins in mouse xenograft tumor tissues treated with DMAPT are also consistent

with the result obtained from in vitro studies (Fig. 3D), indicating that parthenolide decreases the

level of mitochondrial antioxidant proteins in prostate tumors.

Oxidization of Keap1 leads to activation of the Nrf2 pro-survival pathway in normal cells.

Activation of the Nrf2 signaling pathway through dissociation with Keap1 resulting in Nrf2

nuclear translocation is considered to be a primary pro-survival pathway in response to oxidative

stress (30, 37). To examine whether parthenolide changes Nrf2 nuclear translocation, the levels

of Nrf2 in nuclei were measured. As shown in Fig. 4A, the nuclear levels of Nrf2 were increased

in the three normal cell lines treated with parthenolide, but no changes were observed in the three

cancer cell lines. To examine whether activation of the Nrf2 pathway is a major mechanism by

which parthenolide protects normal cells against radiation injury, Keap1 and Nrf2 were silenced

in PZ cells by transfecting their siRNA (Fig. 4B, left panel). Cell survival decreased when Nrf2

was silent. IR significantly reduced cell survival but the cell survival was restored when Keap1

was silenced (Fig. 4B, right panel). These results suggest that oxidation of Keap1 and subsequent

activation of Nrf2 by parthenolide are essential for normal cell survival after radiation treatment.

Thioredoxin is necessary for parthenolide-mediated reduction of Keap1 in cancer cells. TrX

is highly expressed in cancer cells and stimulates cell growth. We previously reported that

parthenolide decreases the reduced form of TrX but increases the oxidized form of TrX in

prostate cancer cells (25). In the present study, we verify that TrX was expressed at a high level

in all three cancer cell lines whereas a low level was observed in the three non-cancer cell lines

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 14: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

14

(Fig. 5A). Immunoprecipitation of Keap1 protein from PC3 cell extracts using a TrX antibody

suggests an interaction between Keap1 and TrX that is increased by parthenolide (Fig. 5B). To

detect whether the parthenolide-influenced reduction of Keap1 in cancer cells is dependent on

TrX, we selectively silenced TrX by transfecting its siRNA prior to parthenolide treatment (Fig.

5C, left panel). As expected, the reduced form of Keap1 was decreased, but the oxidized form of

Keap1 was increased when TrX was silent (Fig. 5C, middle and right panels). The results suggest

that TrX is interacting with Keap1 to keep Keap1 in a reduced state in parthenolide-treated cells.

To further confirm that the function of Keap1 leads to cell death in parthenolide-treated cancer

cells, a Keap1 expression construct was transfected into PC3 cells, followed by parthenolide and

IR treatments. Overexpression of Keap1 resulted in increases in cell death in both treated and

untreated cells (Fig. 5D, top panel). The levels of mitochondrial phosphoglycerate mutase 5

(PGAM5), a protein serine/threonine phosphatase that interacts with Bcl-xL in the mitochondrial

membrane (38), and Bcl-xLwere clearly decreased in the Keap1 transfected cells, but no changes

were observed in Nrf2, Ikkα and IkBα (Fig. 5D, bottom panel). These results suggest that the

parthenolide-increased reduced form of Keap1facilitates Keap1-mediated ubiquitin/proteasome-

dependent degradation of PGAM5 and Bcl-xL, which is an established mechanism for

parthenolide-mediated cell death in cancer cells.

Keap1 triggers PGAM5-mediated Bcl-xL ubiquitin degradation in parthenolide-treated

cancer cells. To further investigate the mechanism by which parthenolide enhances the

radiosensitivity of prostate cancer cells, we determined the interactions between Keap1, PGAM5

and Bcl-xL. The results demonstrate that a reduced form of Keap1, which is increased in

parthenolide-treated PC3 cells, enhanced interaction between Keap1 and PGAM5, as detected by

immunoprecipitation using a PGAM5 antibody (Fig. 6A). Bcl-xL, a prosurvival mitochondrial

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 15: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

15

protein, was also increased in the pulled down complex (Fig. 6A). Interestingly, the proteins that

are associated with Keap1 were decreased in whole cell extracts (Fig. 6B). A time course of

parthenolide treatment shows that PGAM5 and Bcl-xL proteins were slightly increased at 12 h

but decreased at 24 and 48 h after treatment (Fig. 6C). Proteins in different cellular fractions

were also quantified (Fig. 6D). The mitochondria-associated proteins PGAM5 and Bcl-xL were

reduced by the parthenolide treatment, but no change was observed in Hsp75, a control for

mitochondrial protein. Parthenolide had no major effect on the levels of Nrf2 and Ikkα in treated

cells. These results suggest that parthenolide enhances Keap1-mediated ubiquitin/proteasome-

dependent degradation of PGAM5 and Bcl-xL (39). In addition, parthenolide increased the level

of mitochondria-associated autophagic protein LC3B, suggesting that parthenolide may enhance

the radiation sensitivity of prostate cancer cells partially through triggering the autophagy

pathway.

Because Keap1 interacts with PGAM5/Bcl-xL/Nrf2, we decided to determine the effect

of PGAM5/Bcl-xL/Nrf2 in mediating parthenolide’s effect on cancer cells. PGAM5, Bcl-xL and

Nrf2 were silenced using their siRNAs, followed by parthenolide treatment (Fig. 6E, bottom

panel). The cell survival fraction was decreased when PGAM5 or Bcl-xL was silent, which is

similar to the effect of parthenolide (Fig. 6E, top panel). No significant additive effects were

observed when parthenolide was combined with PGAM5 or Bcl-xL siRNA. In contrast, a

significant effect was observed when Nrf2 was silent. These results suggest that Keap1-mediated

PGAM5/Bcl-xL degradation, but not Nrf2 degradation, is important for parthenolide-induced

cancer cell death.

To further determine whether the function of Bcl-xL plays a major role in protecting

cancer cells against parthenolide-induced cell death, a plasmid carrying Bcl-xL cDNA was

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 16: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

16

transfected into PC3 cells followed by parthenolide treatment (Fig. 6F, bottom panel). The

results show that expression of Bcl-xL efficiently protects cells from cytotoxicity caused by

parthenolide (Fig. 6F, top panel). Taken together, these results suggest that parthenolide

enhances the radiosensitivity of prostate cancer cells, in part, by triggering ubiquitin/proteasome-

based degradation of Bcl-xL.

In summary, parthenolide provides radiosensitization in prostate cancer cells but

radioprotection in normal cells, and the observed differential effects are mediated, in part, by

redox modification of Keap1, i.e., reducing Keap1 in cancer cells but oxidizing Keap1 in normal

cells. The distinct redox modification of Keap1 initiates different signaling pathways that affect

mitochondrial function, leading to cell survival or cell injury in response to radiation, as

illustrated in Fig. 7.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 17: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

17

Discussion

The majority of anticancer therapies fail because cancers develop phenotypes that are

treatment resistant and because treatments cause unwanted and/or detrimental side effects to

normal cells or to untargeted tissues. While conventional adjuvant therapies improve tumor

response to radiation therapy, they generally cause additional damage to normal tissues. Thus,

the focus of the present study is to identify adjuvant therapeutics that can reduce the side effects

of radiation therapy. Our study provides a proof-of-concept for improving the efficacy of

radiation therapy while protecting against injury to normal tissues. It has been demonstrated that

parthenolide, the anti-inflammatory phytochemical, is able to suppress tumor growth in many

organs (22-25). In addition, parthenolide appears to synergically enhance chemotherapeutic

efficiency when it is combined with taxol or cisplatin to treat lung and gastric cancer cells (23,

40). Parthenolide also sensitizes radioresistant osteosarcoma cells to radiotherapy (41). Here, we

demonstrate that DMAPT, a parthenolide prodrug, sensitized prostate cancer cells to

radiotherapy in vivo and protected normal prostate and bladder against radiation-induced tissue

injury. These results extend our previous survival studies in prostate cancer cell lines and normal

prostate epithelial cells.

ROS, as products of cell metabolism, play a dual role in tumorigenesis and tumor

suppression. The “two-faced” character of ROS has emerged as a potential source for

discovering anticancer drugs. Redox homeostasis is frequently deregulated in cancers as it is

constantly exposed to high levels of ROS compared to normal counterparts. Our data

demonstrate that constitutively elevated levels of oxidative stress in cancer cells represent a

specific vulnerability that can be selectively targeted by direct- or indirect-acting prooxidants and

antioxidants or redox modulators. Theoretically, the differential redox status of cancer cells

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 18: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

18

compared to normal cells should provide a therapeutic window for selective redox intervention

via additional increases in ROS. In this context, normal and cancer cells should respond

differently to the same level of prooxidant action generated either by direct production of

oxidative species or by modulation of specific cellular targets involved in redox regulation. In

this study we demonstrate that parthenolide serves as a prooxidant and displays a selective redox

modification capability that differentially modulates cellular redox signals and targets. The

Michael acceptor reacts with a thiol group of target proteins through covalent adduction (21).

Parthenolide contains electrophilic α-methylene-γ-lactone, a bisfunctional Michael acceptor, and

displays a potential for bifunctional target alkylation and crosslinking. The present study

demonstrates the inverse effects of parthenolide on redox modification in cancer cells compared

to normal cells. Remarkably, observations of the cytotoxic and cytoprotective effects of

parthenolide are consistent with its action in the modulation of ROS levels in both cancer and

normal cells. Alteration of cellular ROS by parthenolide is attributed to functionally up- or

down-regulating antioxidant enzymes in mitochondria, which consequently regulates

mitochondrial respiration. Parthenolide is able to selectively reduce the activity of several

enzymes involved in oxidative stress removal in cancer cells, which in turn can cause ROS levels

to rise above the threshold for cell death. This finding predicts that antioxidant proteins and

mitochondria are feasible therapeutic targets.

It has been reported that parthenolide is a potent inhibitor of NF-κB, which is a ROS-

responsive transcriptional factor involved in both tumor progression and tumor resistance to

treatment through upregulation of anti-apoptotic genes, such as Bcl-2, Bcl-xL, survivin, and

XIAP (42). We previously demonstrated that NADPH oxidase-mediated inactivation of the Foxo

3 signaling pathway is involved in the parthenolide-enhanced radiosensitivity of prostate cancer

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 19: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

19

(25). However, previous studies did not explain how parthenolide exerts such an opposing effect

in tumor and normal cells. The present study identifies Keap1 as a redox signaling sensor that

plays a pivotal role in the differential regulation of the downstream signaling targets in response

to radiation-mediated cytotoxicity in prostate cancer and normal cells. Keap1, an adaptor protein

for ubiquitin-based processing by the CUL3/RBX1-dependent E3 ubiquitin ligase complex,

functions as a sensor for thiol-reactive redox modification (43). The present study demonstrates

that stabilization of Nrf2 by oxidation of Keap1 serves as a major mechanism by which

parthenolide protects normal tissues against radiotoxicity through up-regulation of antioxidant

enzymes in mitochondria. However, Nrf2 transcriptional activation did not play a major role in

parthenolide-treated prostate cancer cells. Thus, it is interesting to note that unlike traditional

chemotherapeutic agents, parthenolide is unable to enhance resistance of prostate cancer to

radiation treatment by stimulating Nrf2 target genes.

In addition to regulating the Nrf2 signaling pathway, Keap1 is able to bind other proteins

such as p62 and PGAM5 (44). Interaction between Keap1 and p62 facilitates release of Nrf2

from the complex, which is considered to be a noncanonical cysteine-independent mechanism

for the autophagy deficiency-activated Nrf2 pathway (45). The N-terminus of PGAM5 interacts

with the Kelch domain of Keap1 and its C-terminus binds to Bcl-xL. Keap1-dependent

ubiquitination results in proteasome-dependent degradation of PGAM5 and Bcl-xL (38). Bcl-xL,

an important member of the Bcl-2 family, is a potent antiapoptotic factor that plays a crucial role

in cell survival by maintaining the electrochemical and osmotic homeostasis of mitochondria

(46). The present study demonstrates that parthenolide increases the level of reduced Keap1 and

consequently induces Keap1-dependent degradation of PGAM5 and Bcl-xL in cancer cells,

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 20: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

20

suggesting that formation of the Keap1-PGAM5-Bcl-xL complex is a mechanism underlying the

effect of parthenolide on radiosensitization of prostate cancer cells.

Although a high rate of aerobic glycolysis in tumors, known as the Warburg effect, has

been observed in various types of cancer, cancers have functional mitochondria, and

mitochondrial respiration is necessary for cancer cell proliferation (47). Cancer cells depend on a

hyperactive metabolism to fuel their rapid growth and also on antioxidative enzymes to quench

potentially toxic ROS generated by such a high metabolic demand (48). Our results demonstrate

that parthenolide not only suppresses MnSOD and GpX, two major antioxidant enzymes in

mitochondria, but also activates Bcl-xL degradation in cancer cells, which suggests that

mitochondria are a feasible target for anticancer treatment. The present study also shows that

parthenolide may maintain normal cell survival through induction of MnSOD and GpX activity.

Thus, a more efficient and safe therapy may involve modification of cellular redox signaling by

alteration of the antioxidant response coupled to selective degradation of prosurvival members of

the Bcl2 family in cancer cells, because conventional anticancer therapy mainly causes cell

growth arrest or cell death by raising cellular ROS, which oxidizes and damages DNA, proteins

and lipids. Optimizing prototype redox chemotherapeutics from natural sources provides an

exciting opportunity to further develop even better candidates to enhance therapeutic efficacy

with less off-target toxicity.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 21: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

21

Acknowledgments

This work was initiated by a previous graduate student, Yulan Sun, and was supported by

National Institutes of Health grants CA49797, CA115801, and CA143428 to Daret K. St. Clair

and William St. Clair. Additional support was provided by the Edward P. Evans Foundation and

by the resources and facilities of the William S. Middleton Veterans Administration Hospital

(Madison, WI).

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 22: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

22

References

1. Irani K, Xia Y, Zweier JL, Sollott SJ, Der CJ, Fearon ER, et al. Mitogenic signaling mediated

by oxidants in Ras-transformed fibroblasts. Science. 1997; 275:1649-52.

2. Rhee SG. Cell signaling. H2O2, a necessary evil for cell signaling. Science 2006; 312:1882-

83.

3. Pelicano H, Carney D, Huang P. ROS stress in cancer cells and therapeutic implications.

Drug Resist Updat 2004;7:97-110.

4. Tew KD, Townsend DM. Redox platforms in cancer drug discovery and development. Curr

Opin Chem Biol. 2011;15:156-61.

5. D’Autreaux B, Toledano MB. ROS as signalling molecules: Mechanisms that generate

specificity in ROS homeostasis. Nat Rev Mol Cell Biol 2007; 8:813-24.

6. Xia C, Meng Q, Liu LZ, Rojanasakul Y, Rojanasakul Y, Wang XR, Jiang BH. Reactive

oxygen species regulate angiogenesis and tumor growth through vascular endothelial growth

factor. Cancer Res 2007;67:10823-30.

7. Dröge, W. (2002). Free radicals in the physiological control of cell function. Physiol Rev

2002;82:47-95.

8. Valko M, Leibfritz D, Moncol J, Cronin MT, Mazur M, Telser J. Free radicals and

antioxidants in normal physiological functions and human disease. Int J Biochem Cell Biol

2007; 39:44-84.

9. Cabello CM, Bair WB 3rd, and Wondrak GT. Experimental therapeutics: Targeting the redox

Achilles heel of cancer. Curr Opin Investig Drugs 2007; 8:1022-37.

10. Trachootham D, Lu W, Ogasawara MA, Nilsa RD, Huang P. Redox regulation of cell

survival. Antioxid Redox Signal 2008;10:1343-74.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 23: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

23

11. Wang J, Yi J. Cancer cell killing via ROS to increase or decrease, that is the question. Cancer

Biol Ther 2008;7:1875-84.

12. Spyratou E, Makropoulou M, Mourelatou EA, Demetzos C. Biophotonic techniques for

manipulation and characterization of drug delivery nanosystems in cancer therapy. Cancer

Lett 2012 Jan 17.

13. Oberley LW, Oberley TD. Role of antioxidant enzymes in cell immortalization and

transformation. Mol Cell Biochem. 1988; 84:147-53.

14. Zhao Y, Chaiswing L, Oberley TD, Batinic-Haberle I, St Clair W, Epstein CJ, et al. A

mechanism-based antioxidant approach for the reduction of skin carcinogenesis. Cancer Res

2005; 65:1401-5.

15. Dhar SK, Tangpong J, Chaiswing L, Oberley TD, St Clair DK. Manganese superoxide

dismutase is a p53-regulated gene that switches cancers between early and advanced stages.

Cancer Res 2011; 71:6684-95.

16. Josson S, Xu Y, Fang F, Dhar SK, St Clair DK, St Clair WH. RelB regulates manganese

superoxide dismutase gene and resistance to ionizing radiation of prostate cancer cells.

Oncogene 2006; 25:1554-59.

17. Fan M, Ahmed KM, Coleman MC, Spitz DR, Li JJ. Nuclear factor-kappaB and manganese

superoxide dismutase mediate adaptive radioresistance in low-dose irradiated mouse skin

epithelial cells. Cancer Res 2007; 67:3220-28.

18. Homma S, Ishii Y, Morishima Y, Yamadori T, Matsuno Y, Haraguchi N, et al. Nrf2

enhances cell proliferation and resistance to anticancer drugs in human lung cancer. Clin

Cancer Res 2009; 15:3423-32.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 24: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

24

19. Hassane DC, Sen S, Minhajuddin M, Rossi RM, Corbett CA, Balys M, et al. Chemical

genomic screening reveals synergism between parthenolide and inhibitors of the PI-3 kinase

and mTOR pathways. Blood. 2010; 116:5983-90.

20. Raj L, Ide T, Gurkar AU, Foley M, Schenone M, Li X, et al. Selective killing of cancer cells

by a small molecule targeting the stress response to ROS. Nature 2011; 475:231-34.

21. Hwang DR, Wu YS, Chang CW, Chang CW, Lien TW, Chen WC, et al. Synthesis and anti-

viral activity of a series of sesquiterpene lactones and analogues in the subgenomic HCV

replicon system. Bioorg Med Chem 2006; 14:83-91.

22. Pajak B, Gajkowska B, Orzechowski A. Molecular basis of parthenolide-dependent

proapoptotic activity in cancer cells. Folia Histochem Cytobiol 2008; 46:129-35.

23. Gill KK, Kaddoumi A, Nazzal S. Mixed micelles of PEG(2000)-DSPE and vitamin-E TPGS

for concurrent delivery of paclitaxel and parthenolide: enhanced chemosensitization and

antitumor efficacy against non-small cell lung cancer (NSCLC) cell lines. Eur J Pharm Sci.

2012; 46:64-71.

24. Sun Y, St Clair DK, Fang F, Warren GW, Rangnekar VM, Crooks PA, et al. The

radiosensitization effect of parthenolide in prostate cancer cells is mediated by nuclear factor-

kappaB inhibition and enhanced by the presence of PTEN. Mol Cancer Ther. 2007;9:2477-

86.

25. Sun Y, St Clair DK, Xu Y, Crooks PA, St Clair WH. A NADPH oxidase-dependent redox

signaling pathway mediates the selective radiosensitization effect of parthenolide in prostate

cancer cells. Cancer Res 2010; 70:2880-90.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 25: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

25

26. Kishida Y, Yoshikawa H, Myoui A. Parthenolide, a natural inhibitor of Nuclear Factor-

kappaB, inhibits lung colonization of murine osteosarcoma cells. Clin Cancer Res 2007;

13:59-67.

27. Salmeen A, Andersen JN, Myers MP, Meng TC, Hinks JA, Tonks NK, et al. Redox

regulation of protein tyrosine phosphatase 1B involves a sulphenyl-amide intermediate.

Nature 2003; 423:769-73.

28. Rasmussen HH, Hamilton EJ, Liu CC, Figtree GA. Reversible oxidative modification:

implications for cardiovascular physiology and pathophysiology. Trends Cardiovasc Med

2010; 20:85-90.

29. Wakabayashi N, Dinkova-Kostova AT, Holtzclaw WD, Kang MI, Kobayashi A, Yamamoto

M, et al. Protection against electrophile and oxidant stress by induction of the phase 2

response: fate of cysteines of the Keap1 sensor modified by inducers. Proc Natl Acad Sci

USA 2004;101:2040-45.

30. Adam J, Hatipoglu E, O'Flaherty L, Ternette N, Sahgal N, Lockstone H, et al. Renal cyst

formation in Fh1-deficient mice is independent of the Hif/Phd pathway: roles for fumarate in

KEAP1 succination and Nrf2 signaling. Cancer Cell 2011; 20:524-37.

31. Neelakantan S, Nasim S, Guzman ML, Jordan CT, Crooks PA. Aminoparthenolides as novel

anti-leukemic agents: Discovery of the NF-kappaB inhibitor, DMAPT (LC-1). Bioorg Med

Chem Lett 2009;19:4346-9.

32. Xu Y, Fang F, St Clair DK, Josson S, Sompol P, Spasojevic I, et al. Suppression of RelB-

mediated manganese superoxide dismutase expression reveals a primary mechanism for

radiosensitization effect of 1alpha, 25-dihydroxyvitamin D(3) in prostate cancer cells. Mol

Cancer Ther 2007; 6:2048-56.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 26: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

26

33. Xu Y, Josson S, Fang F, St Clair DK, Wan XS, Sun Y, et al. RelB enhances prostate cancer

growth: implications for the role of the NF-κB alternative pathway in tumorigenicity. Cancer

Res 2009; 69:3267-71.

34. Slane BG, Aykin-Burns N, Smith BJ, Kalen AL, Goswami PC, Domann FE, et al. Mutation

of succinate dehydrogenase subunit C results in increased O2.-, oxidative stress, and genomic

instability. Cancer Res 2006; 66:7615-20.

35. Bayer EA, Safars M, Wilchek M. Selective labeling of sulfhydryls and disulfides on blot

transfers using avidin-biotin technology: studies on purified proteins and erythrocyte

membranes. Anal Biochem 1987; 161:262-71.

36. Keller JN, Kindy MS, Holtsberg FW, St Clair DK, Yen HC, Germeyer A, et al.,

Mitochondrial manganese superoxide dismutase prevents neural apoptosis and reduces

ischemic brain injury: suppression of peroxynitrite production, lipid peroxidation, and

mitochondrial dysfunction. J Neurosci 1998;18:687-97.

37. Wu RP, Hayashi T, Cottam HB, Jin G, Yao S, Wu CC, et al. Nrf2 responses and the

therapeutic selectivity of electrophilic compounds in chronic lymphocytic leukemia. Proc

Natl Acad Sci U S A 2010;107:7479-84.

38. Lo SC, Hannink M. PGAM5, a Bcl-xL-interacting protein, is a novel substrate for the redox-

regulated Keap1-dependent ubiquitin ligase complex. J Biol Chem 2006; 281:37893-903.

39. Niture SK, Jaiswal AK. Inhibitor of Nrf2 (INrf2 or Keap1) protein degrades Bcl-xL via

phosphoglycerate mutase 5 and controls cellular apoptosis. J Biol Chem 2011; 286:44542-56.

40. Sohma I, Fujiwara Y, Sugita Y, Yoshioka A, Shirakawa M, Moon JH, et al. Parthenolide, an

NF-κB inhibitor, suppresses tumor growth and enhances response to chemotherapy in gastric

cancer. Cancer Genomics Proteomics 2011; 8:39-47.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 27: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

27

41. Zuch D, Giang AH, Shapovalov Y, Schwarz E, Rosier R, O'Keefe R, et al. Targeting

radioresistant osteosarcoma cells with parthenolide. J Cell Biochem 2012; 113:1282-91.

42. Montero AJ, Jassem J. Cellular redox pathways as a therapeutic target in the treatment of

cancer. Drugs 2011; 71:1385-96.

43. Sekhar KR, Rachakonda G, Freeman ML. Cysteine-based regulation of the CUL3 adaptor

protein Keap1. Toxicol Appl Pharmacol 2010; 244:21-26.

44. Lau A, Wang XJ, Zhao F, Villeneuve NF, Wu T, Jiang T, et al. A noncanonical mechanism

of Nrf2 activation by autophagy deficiency: direct interaction between Keap1 and p62. Mol

Cell Biol. 2010; 30:3275-85.

45. Rusten TE, Stenmark H. p62, an autophagy hero or culprit? Nat Cell Biol 2010; 12:207-9.

46. Vander Heiden MG, Chandel NS, Williamson EK, Schumacker PT, Thompson CB. Bcl-xL

regulates the membrane potential and volume homeostasis of mitochondria. Cell 1997;

91:627-37.

47. Weinberg F, Chandel NS. Mitochondrial metabolism and cancer. Ann NY Acad Sci 2009;

1177:66-73.

48. Stephanos Pavlides, Iset Vera, Ricardo Gandara, et al. Warburg Meets Autophagy: Cancer-

Associated Fibroblasts Accelerate Tumor Growth and Metastasis via Oxidative Stress,

Mitophagy, and Aerobic Glycolysis. Antioxidants & Redox Signaling 2012;1264-84.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 28: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

28

Figure Legends Figure 1. The effect of parthenolide on radiosensitivities of prostate cancer and normal

cells. (A) Prostate cancer PC3 cells were injected into the flanks of nude male mice. The

resulting tumors were treated with DMAPT and IR. Tumor volume was measured and tumor

growth was calculated. (B) Time needed for tumor growth to reach 2000 mm3 volume after

treatment was calculated and plotted. (C) Mice without cancer were treated with radiation alone

(5 x 3 Gy) and DMAPT (10 mg/kg) with radiation. Prostate and bladder tissues were removed

for pathological analysis using EM. Asterisks indicate normal mitochondria and arrows indicate

mitochondria with myelin figures. M, normal mitochondria; Ly, lysosome; and V, mitochondria

with vacuoles. (D) Quantification of mitochondrial damage in mice prostate and bladder tissues.

Figure 2. The effect of parthenolide on redox homeostasis in prostate cancer and normal

cells. (A-B) Cells were treated with parthenolide and then labeled with DHE or DCF. The mean

fluorescence intensity of DHE (A) and the ratio of H2DCFDA to DCFDA (B) were determined

using flow cytometry. Concentrations of cellular superoxide and total ROS were estimated by

quantification of fluorescence intensity. Antimycin and PMA were used as positive controls for

generation of ROS. PEG-SOD was used as a control to remove superoxide generated by DHE.

PN, parthenolide. (C) The levels of antioxidant proteins were quantified by western blots. (D)

The activities of the corresponding enzymes were measured. (E) Quantification of the basal

oxygen consumption, ATP-linked oxygen consumption, the maximal OCR after the addition of

FCCP, and the reserve capacity of the cells. NS, not significant. (F) Three cancer cell lines and

three non-cancer cell lines were treated with parthenolide (PN) at the indicated concentrations.

Cell survival fraction was determined by MTT assay.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 29: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

29

Figure 3. Keap1 redox modification by parthenolide. (A-B) Three normal cell lines (A) and

three prostate cancer cell lines (B) were treated with saline or parthenolide (PN). Keap1 was

immunoprecipitated with its antibody and SH- and S-S moieties of Keap1 were demonstrated by

ECL detection. (C) Tumor tissues were homogenized and incubated with MPB. Reduced (SH-)

and oxidized (S-S) forms of Keap1 in the tissues were detected. (D) Tumor tissues were

homogenized and antioxidant proteins were quantified by western blots. The right panel shows

representative blots and the left panel shows the average of multiple blots.

Figure 4. Activation of Keap1-Nrf2 pathway by parthenolide in normal cells. (A)

Parthenolide (PN) increases nuclear levels of Nrf2 in normal cells but not in cancer cells.

Nuclear proteins extracted from the parthenolide-treated cell lines were immunoblotted to

quantify nuclear levels of Nrf2 using PCNA as loading control. (B) The effect of Keap1 and

Nrf2 in normal cells. PZ cells were transfected with siRNAs to knock down Keap1 or Nrf2,

respectively. After treatment with parthenolide (PN) and IR, the cell survival fraction was

quantified using Trypan blue exclusion assay (right), and the knocked-down Keap1 and Nrf2

were confirmed by western blots (left).

Figure 5. TrX-dependent Keap1 reduction by parthenolide in prostate cancer cells. (A) The

levels of TrX in normal and cancer cells before and after treatment with parthenolide (PN) were

detected by western blots. (B) After the indicated treatment, Keap1 was immunoprecipitated

using a TrX antibody. (C) PC3 cells were transfected with TrX siRNA prior to the indicated

treatment (top). SH- and S-S bands in Keap1 protein were detected as described in Fig. 3

(bottom). (D) A Keap1 cDNA construct was transfected into PC3 cells. The levels of related

proteins were detected by western blots (bottom). The effect of Keap1 on radiosensitivity was

analyzed using a Trypan blue exclusion assay (top).

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 30: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

30

Figure 6. Degradation of PGAM5-Bcl-xL caused by parthenolide-mediated reduction of

Keap1 in prostate cancer cells. (A) PC3 cells were treated with parthenolide (PN). Keap1 and

Bcl-xL were immunoprecipitated using a PGAM5 antibody. (B) The total levels of Nrf2,

PGAM5, and Bcl-xL were quantified by western blots. (C) The levels of mitochondria-

associated proteins after parthenolide treatments. (D) Proteins in various cellular fractions were

identified with antibody specific for each protein. (E) PC3 cells were transfected with siRNA to

knock down PGAM5, Bcl-xL and Nrf2, respectively (bottom). Cell survival fraction was

quantified by Trypan blue exclusion assay (top). (F) A Bcl-xL expression construct was

transfected into PC3 cells and the expression of Bcl-xL was monitored by western blots

(bottom). Cell survival fraction was quantified by Trypan blue exclusion assay (top).

Figure 7. A proposed mechanistic model for parthenolide-mediated inverse therapeutic

effects on radiosensitivity of prostate cancer and radioresistance of normal cells.

Parthenolide sensitizes cancer cells to radiation, in part, by maintaining Keap1 in a reduced state

and enhancing its interaction with PGAM5 and Bcl-xL, resulting in degradation of Bcl-xL in

mitochondria. In contrast, parthenolide protects normal cells against radiation via oxidation of

Keap1 and release of the Nrf2 transcription factor for activation of mitochondrial antioxidant

enzymes.

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297

Page 31: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 1

A CA5 x 3 Gy DMAPT + 5 x 3 Gy

Prostate

C

a b

me

(mm

3 )

2000

2500

3000untreated

DMAPT

1 Gy

1 Gy + DMAPT

2 Gy

2 Gy + DMAPT

c d

12,000X 12,000X

Tum

or v

olum

500

1000

1500y

B

Bladder

D

8,000X 8,000X

p<0.05

Day after treatment

00 2 4 6 8 11 14 17 20 24 28 32 36

60

80

1005x3Gy DMAPT/5x3Gy

*

ondr

ia /

Tota

l dr

ia

* p<0.05

30

40

50

60

70

p<0.05

p<0.01

tum

or v

olum

e20

00 m

m3

20

40 *

Dam

aged

mito

cho

mito

chon

d0

10

20

30

Day

s w

hen

reac

hes

to

0

Prostate Bladder

% D

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 32: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 2

Aity

DHE assay 2000

PC3

PZ

P EC* 160

180

200 PC3/DMSO

PC3/PN

PZ/DMSO

* P<0.05

mg)

D

n flu

ores

cenc

e in

tens

500

1000

1500PrEC

*#

$

60

80

100

120

140 PZ/PN

PrEC/DMSO

PrEC/PN

*

*

yme

activ

ity (u

nits

/min

/m

B

Mea

PZP 0 01

0DMSO PN (5μM) PN+PEG-SOD

(50 u/ml)Antimycin

(5 μM)

DCF assay 120 PC3

E

0

20

40

MnSOD CuZnSOD catalase GpX

*

*

*

*

Enzy

150

200

250

300

es/m

in/m

gpro

tein

)

PZPZ+PNPC3PC3+PN

NS

P<0.01

P<0.05

P<0.01

P<0.0140

60

80

100PZPrEC* *

# $

E

H2D

CFD

A/D

CFD

A

cancer cells normal cells 0

50

100

Basal ATP-Linked Maximal OCR ReserveCapacity

OC

R(p

Mol NS

NS NS

0

20

DMSO PN (5 μM) PMA (5 μM)

C

H

MnSOD

PN (5 μM) - + - + - + - + - + - +

GpX

Fll

surv

ival

(%)

60

80

100

120PC3DU-145LNCaPPZRWPE-1PrEC

β-actin

CuZnSOD

catalase

Ce

0

20

40

0 1 2 3 4 5

PrEC

PN (μM)

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 33: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 3

APrECPZ-HPV-7 RWPE-1 -SH -S-S-

C

PN - + - +

Keap1

PN - + - +

Keap1

PN - + - +

Keap1

-SH-S-S- -SH-S-S- -SH-S-S-

-SH -S-S-

DMAPT - + - +

Keap1p p p

B

IgG IgG IgG

-SH-S-S-

PC-3

-SH-S-S-

DU-145

-SH-S-S-

LNCaP

IgG

DMSO

D

SH

PN - + - +

Keap1

S S SH

PN - + - +

Keap1

S S SH

PN - + - +

Keap1

S S

e (%

)

* p<0.05

80

100

120

140

DMSO

DMAPT

*

MnSOD

GPX

Fold

cha

nge

0

20

40

60

80 * *

catalase

GPX

CuZnSOD

β-actin

IgG IgG IgG MnSOD GpX CuZnSOD catalase

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 34: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 4

A

PN (5 μM) - + - + - + - + - + - +

PC3 DU-145 LNCaP PZ PrEC RWPE-1

Nrf2

B

PCNA

B

P<0.01

P<0.01

140 ctrl siRNA

Keap1

PN (5 μM) - + - + - +

urvi

val (

%)

60

80

100

120 Keap1 siRNA

Nrf2 siRNA

Nrf2

β-actin

Cel

l su

0

20

40

Untreated 6Gy PN 5uM PN + 6Gy

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 35: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 5

PN (5 μM) - + +

IP TrX IgGA

B

β

TrX

- + - + - + - + - + - +

PC3 DU145 LNCaP PZ PrEC RWPE -1

PN (5μM)

PN (5 μM) + +

Keap1

β-actin

TrXC

siRNA ctrl siRNA TrX

D

Keap1

TrX

PN (5μM) - + - +

60

80

100

120

VectorKeap1

*

*

* p<0.05

surv

ival

(%)

D

siRNA ctrl siRNA TrX

-SH -S-S-

β-actin

siRNA ctrl siRNA TrX

Vector Keap1

0

20

40

Untreated 6Gy PN(5μM) PN+6Gy

**Ce

ll s

PN (5μM) - + - +

Keap1

- + - +

PGAM5

Nrf2

Keap1

PN (5μM)- + - +

Vector Keap1

IgGβ-actin

Bcl-xL

Ikkα

IkBα

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 36: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 6

A IP: PGAM5

I G

D

E

Keap1

PN(5μM)

PGAM5

Bcl xL

+ - +

IgG

al (%

)

p<0.01

p<0.01

80

100

120DMSOPN

p<0.01

p<0.01PGAM5

Keap1

Nrf2

- + - + - +

BclxL

PN(5μM)

Keap1

PN(5μM) - +

Bcl-xL

IgG

B

Cell

surv

iva

0

20

40

60LC3B1LC3B2

Ikkα

Hsp75

Bcl-xL

Bcl-xL

Nrf2

p

PGAM5

0ctrl siRNA PGAM5 siRNA Bcl-xL siRNA Nrf2 siRNAβ-actin

120

DMSO

PN

p<0.01

F

LC3B1

LC3B2

β-actin

C

Nrf2

PGAM5

B l L

PN (5μM) - + - + - + - +

40

60

80

100

Cell

surv

ival

(%)

PN(5μM) - + - + - +

12h 24h 48h

Keap1

PGAM5

Bcl-xL

β-actin0

20

vector Bcl-xL

BcL-xL

β-actin

PN (5μM) - + - +

Bcl-xL

β-actin

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 37: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Figure 7

on April 15, 2021. ©

2013 Am

erican Association for C

ancer Research.

cancerres.aacrjournals.org D

ownloaded from

Author m

anuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Author M

anuscript Published O

nlineFirst on M

ay 14, 2013; DO

I: 10.1158/0008-5472.CA

N-12-4297

Page 38: The authors disclose no potential conflicts of interestcancerres.aacrjournals.org/content/canres/... · 5/14/2013  · The tumors were treated five times with 10 mg/kg DMAPT and 3

Published OnlineFirst May 14, 2013.Cancer Res   Yong Xu, Fang Fang, Sumitra Miriyala, et al.   Cancer CellsOpposing Radiosensitive Effects of Parthenolide in Normal and KEAP1 Is a Redox Sensitive Target That Arbitrates the

  Updated version

  10.1158/0008-5472.CAN-12-4297doi:

Access the most recent version of this article at:

  Manuscript

Authoredited. Author manuscripts have been peer reviewed and accepted for publication but have not yet been

   

   

   

  E-mail alerts related to this article or journal.Sign up to receive free email-alerts

  Subscriptions

Reprints and

  [email protected] at

To order reprints of this article or to subscribe to the journal, contact the AACR Publications

  Permissions

  Rightslink site. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC)

.http://cancerres.aacrjournals.org/content/early/2013/05/14/0008-5472.CAN-12-4297To request permission to re-use all or part of this article, use this link

on April 15, 2021. © 2013 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on May 14, 2013; DOI: 10.1158/0008-5472.CAN-12-4297